YOU ARE DOWNLOADING DOCUMENT

Please tick the box to continue:

Transcript
Page 1: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

9374 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 This journal is c the Owner Societies 2011

Cite this: Phys. Chem. Chem. Phys., 2011, 13, 9374–9384

Molecular dynamic simulation of dicarboxylic acid coated aqueous

aerosol: structure and processing of water vapor

Xiaofei Ma,aPurnendu Chakraborty,

aBrian J. Henz

cand Michael R. Zachariah*

ab

Received 23rd September 2010, Accepted 15th March 2011

DOI: 10.1039/c0cp01923b

Organic monolayers at the surfaces of aqueous aerosols play an important role in determining the

mass, heat transfer rate and surface reactivity of atmospheric aerosols. They can potentially

contribute to the formation of cloud condensation nuclei (CCN) and are involved in a series

of chemical reactions occurring in atmosphere. Recent studies even suggest that organic-coated

interfaces could have played some role in prebiotic biochemistry and the origin of life. However,

creating reproducible, well-characterized aqueous aerosol particles coated with organic films is

an experimental challenge. This opens the opportunity for computer simulations and modeling

of these complex structures. In this work, molecular dynamics simulation was used to probe

the structure and the interfacial properties of the dicarboxylic acid coated aqueous aerosol.

Low molecular weight dicarboxylic acids of various chain lengths and water solubility were

chosen to coat a water droplet consisting of 2440 water molecules. For malonic acid coated

aerosol, the surface acid molecules dissolved into the water core and formed an ordered structure

due to the hydrophobic interactions. The acid and the water are separated inside the aerosol.

For other nanoaerosols coated with low solubility acids, phase separation between water and

acid molecules was observed on the surface of the particle. To study the water processing

of the coated aerosols, the water vapor accommodation factors were calculated.

Introduction

Organic material is ubiquitous in the earth’s atmosphere and

represents an important fraction of the fine aerosol mass.

Studies have shown that total organic carbon can represent

10–65% of the aerosol mass and exists as a complex mixture of

hundreds of organic compounds, while secondary organic

carbon can contribute up to 25–50% of the fine aerosol mass

in urban polluted areas.1 Indirectly, atmospheric aerosols can

affect the radiative properties and lifetime of clouds and thus

have an influence on global climate by acting as cloud

condensation nuclei (CCN).2 Observations have revealed that

more than 60% of the CCN can consist of organic constituents.3

Recent experimental studies and thermodynamic analysis of

organic marine aerosols even suggest that atmospheric aerosols

could act as prebiotic chemical reactors and play a role in the

origin of life.4,5 Despite the considerable fraction of organic

matters in atmospheric aerosols and significant importance of

their environmental and biological functions, little is known

about their structure and influence on atmospheric processes.

The organic materials can be water-soluble and insoluble,

volatile and nonvolatile, surface-active and surface-inactive,

and biogenic and anthropogenic. Depending on their physical

properties (e.g. solubility and volatility), the organics can form

different structured films on existing aerosol particle surfaces.

Water-insoluble organic molecules are likely to be closed-packed

and oriented and thus tend to form ‘‘condensed films’’ on the

particle surfaces. Phase transitions which correspond to differing

degrees of ordering of the surfactant molecules can take place

in those films.6 Our previous molecular dynamic simulation

results7,8 on the structure of long-chain fatty acid coated

nanoaerosols showed that in the final stage of equilibrium,

an inverted micelle structure is formed, and consistent with

a previously proposed ‘‘inverted micelle’’ model.9 In this

structure a water core is surrounded by surface adsorbed fatty

acid molecules. On the other hand, water-soluble organic

surfactant molecules tend to form less compact films which

do not undergo phase transitions to more compact structures.

The primary relevance of these structures is how they

subsequently interact with other organics, accommodation of

water vapor, and its ability to act as cloud condensation nuclei.

There is experimental evidence that organic compounds perturb

the uptake of trace gases onto aqueous surfaces.10 The presence

of the organic films on water drops could significantly alter both

condensation and evaporation rates. To complicate this already

difficult problem, atmospheric ‘‘processing’’ of the surface by

aDepartment of Mechanical Engineering and Department ofChemistry and Biochemistry, University of Maryland, College Park,MD 20742, USA. E-mail: [email protected]

bNational Institute of Standards and Technology, Gaithersburg,MD 20899, USA

cU.S. Army Research Laboratory, Aberdeen Proving Ground,MD 21005, USA

PCCP Dynamic Article Links

www.rsc.org/pccp PAPER

Page 2: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 9375

atmospheric oxidants would further alter the surface properties

of the aerosol, leading to further changes in reactivity.6,9

Among the various kinds of organic compounds, low

molecular weight dicarboxylic acids have attracted much

attention due to their large prevalence and interesting physico-

chemical properties. These types of acids have been identified

as one of the major organics in both urban and rural areas and

are a ubiquitous organic aerosol constituent in the marine and

even Arctic atmosphere.11,12 Observations have shown that

dicarboxylic acids are also commonly found in the organic

fraction of secondary aerosols. However, their formation and

partition to the aerosol phase are still unclear. In biology,

dicarboxylic acids are important metabolic products of fatty

acids. During recent years, a considerable effort has been

made to understand the properties of low molecular weight

dicarboxylic acids (C3–C9). It is known that the physico-

chemical properties of low molecular weight dicarboxylic acids

such as solubility, vapor pressure, evaporation rate, melting

and boiling points alternate with the number of carbon

atoms.13,14 Those physicochemical properties have profound

effect on the CCN activity. One of the major questions

surrounding organic compounds focuses on the changes in

surface tension of the droplet due to the presence of the

organic and the solubility of the compound.2 Water-soluble

materials are known to affect droplet activation by lowering

the surface tension and thus changing the critical droplet

radius. Experiments have confirmed that this effect can be

well predicted by the Kohler theory for soluble inorganic

species and organics that are wettable by water,15 when the

two components are homogenously dispersed. However,

when considering extending the current theory, low-solubility

organic species are equally important. The low molecular

weight dicarboxylic acids (C3–C9) cover a wide range of

solubility and thus provide an excellent platform to study

the effect of solubility on the CCN activation.

With hydrophilic groups at both ends of a hydrophobic

hydrocarbon chain, dicarboxylic acids are bolaamphiphilic

molecules. The structure and phase behavior of these

molecules in a particular type of medium are determined by

unique intermolecular interactions: the hydrophobic inter-

actions between hydrocarbon chains, the hydrophilic and/or

the electrostatic interactions between the head groups. When

amphiphilic molecules are dispersed in water, the hydrophobic

interactions of the hydrocarbon chains drive the molecules to

self-assemble into structures where the hydrophobic tails are

shielded from unfavorable interactions with water by the

hydrophilic, polar head groups.16 Like amphiphilic molecules,

bolaamphiphilic molecule aggregation is driven by hydrophobic

interactions, which can form self-assembly structures, including,

spherical lipid particles, vesicles produced from long-chain

molecules, and micelles from short-chain, water-soluble

bolaamphiphiles.17 Not surprisingly compared with amphilic

molecules, the introduction of a second hydrophilic head

group generally induces a higher solubility in water and an

increase in the critical micelle concentration.

Previous experiments have been carried out to investigate

the cloud activity of various pure dicarboxylic acid aerosols

from highly-soluble acids to almost insoluble acids.11,15,18 The

TDMA (Tandem DMA) method is frequently used in

laboratory studies. In this method, the first DMA produces

nearly monodisperse particles of a known size while the second

DMA measures the particle size distribution of the final

aerosol. However, there are some drawbacks associated with

this method. First, the DMAs are designed for classifying

spherical particles, and the results are interpreted based on a

singly charged assumption. Therefore, depending on particle

morphology, mass and cross-sectional area which affects the

charging efficiency, there is a possibility that DMA could

lead to incorrect size classification. Second, the ion-mobility

method deals with the whole aerosol population, so it is

impossible to use this method to monitor the water processing

of individual particles. In fact, since molecular processes

involve dynamics happening over short distances (nanometer

length scale) and short times (nanoseconds time scale), these

processes are difficult to probe experimentally. Furthermore,

the atmospheric organic aerosols usually consist of more than

one chemical species, and the organics can form complex

structures such as monolayers, thin films on the aerosol

surfaces. It is an experimental challenge to create reproducible,

well-characterized aqueous aerosol particles which are coated

with an organic film. However, on the other hand, the

structure of complex aerosols can be explicitly defined in

molecular simulations and the dynamics of the molecular-

scale process can be followed explicitly as well.

In this work, we employed a molecular dynamic simulation

method to study the structure evolution and water processing

of various dicarboxylic acids coated water droplets. Two

questions are addressed in our study: (1) What is the relation-

ship between the aerosol final structure and physical properties

of coated dicarboxylic acid molecules? (2) How does the

aerosol structure affect the water processing of the coated

aerosol?

Computational model and simulation details

The molecular dynamics simulations in this work were carried

out using Lammps19 (Large-scale Atomic/Molecular Massively

Parallel Simulator, http://lammps.sandia.gov) a software

package developed by Sandia National Laboratories.

1 Simulating the water core

For simulating the water molecule, we used the extended

simple point charge (SPC/E) interaction potential developed

by Berendsen et al.20,21 This model consists of a total of three

sites for the electrostatic interactions of water with an OH

distance of 0.1 nm and HOH angle of 109.471. The partial

positive charges of +0.4328e on the hydrogen atoms are

exactly balanced by partial negative charges of �0.8476elocated on the oxygen atom. A pairwise Lennard-Jones inter-

action centered on the oxygen atoms is used to compute

the van der Waals forces between two water molecules. The

expression is given by:

VLJðrijÞ ¼ �A

rij

� �6

þ B

rij

� �12

ð1Þ

withA=0.37122 (kJmol�1)1/6 nm,B=0.3428 (kJmol�1)1/12 nm,

where rij is the distance between two oxygen atoms. The SPC/E

Page 3: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

9376 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 This journal is c the Owner Societies 2011

potential has been studied extensively. It can provide accurate

reproduction of water surface tension and its temperature

dependence22 and the effects of simulation size and the treat-

ment of long-range interactions on surface tension have been

shown to be less than 10%.23

The first step toward building a dicarboxylic acid coated

aqueous aerosol was to build a pure water droplet.

The detailed procedures for preparing an equilibrated water

droplet are described in our previous work.7 In brief, the water

droplet is evolved from a simple cubic lattice structure with

the oxygen atom at the vertex of each cube. The initial

sphere configuration consisting of 2440 water molecules was

generated by considering only the water molecules inside a

sphere of a certain radius. The water droplet is equilibrated at

the temperature of 260 K. The SHAKE algorithm24 was used

to apply constraints between the O–H bond and H–O–H angle

to maintain the rigidity of the water molecules. The radial

distribution function confirms the liquid phase of the prepared

water droplet.

2 Modeling the dicarboxylic acid molecule

The next step was to coat the spherical water droplet with

dicarboxylic acid molecules. By identifying the surface molecules

of the water droplet, we placed dicarboxylic acid molecules on

the surface water sites with one of the carboxyl groups of the

dicarboxylic acid attached to the surface water molecule and the

straight chain of the dicarboxylic acid placed radially outward.

The initial coating density is about 6.6 molecules per nm2.

The dicarboxylic acids simulated were malonic acid (C3),

succinic acid (C4), glutaric acid (C5), adipic acid (C6), pimelic

acid (C7), suberic acid (C8), azelaic acid (C9) and branched

azelaic acid (C9_branched). The physical properties of these

dicarboxylic acids are summarized in Table 1. For modeling

the dicarboxylic acid molecule, we used a mixture of the

‘‘fully atomistic’’ method and the ‘‘united atom’’ method.

The high-accuracy/high computational requirement fully

atomistic method is used to model the carboxyl group

(–COOH group) while the less accurate/lower computational

requirement united atom method is employed to simulate the

methylene group (–CH2 group). In the united atom setup, the

number of interaction sites is reduced by subsuming some or

all of the atoms into the atoms to which they are bonded.

Thus considerable computation savings are possible. In our

simulation, each methylene group (–CH2 group) is represented

by a single site with interactions defined between these sites.

An example of the dicarboxylic acid in this work is shown in

Fig. 1. During the simulation, the O–H bond in the carboxyl

group was kept rigid using the SHAKE algorithm.

The reliability of predictions from molecular simulations is

determined largely by the accuracy of the representation of the

intermolecular interaction potentials. The potentials used in

our model include nonbonded interactions between each pair

of atoms and bonded interactions between bonded atoms

which have contributions from bond stretching, angle vibra-

tion, proper and improper dihedral interactions. For non-

bonded force calculation, the switched Lennard-Jones and

the switched coulombic potential were used to simulate the

van der Waals interactions and the electrostatic interactions,

respectively. The switched forces are given by:

E ¼

LJðrÞ rorin

SðrÞ � LJðrÞ rinororout

0 r4rout

8>>><>>>:

E ¼

CðrÞ rorin

SðrÞ � CðrÞ rinororout

0 r4rout

8>>><>>>:

LJðrÞ ¼ 4esr

�12� s

r

�� 6� ��

CðrÞ ¼ Cqiqj

er

SðrÞ ¼ ½r2out � r2�2½r2out þ 2r2 � 3r2in�

½r2out � r2in�3

ð2Þ

Harmonic potentials of the form E= K(x � x0)2 (where K is a

prefactor, x is the position vector or angle and x0 is the

corresponding equilibrium value) were chosen to model

bond stretching and angle vibrations. The periodic function

E = K[1 + cos(nf � d)] was used to describe the proper

dihedral interactions. For the improper interactions, a harmonic

potential E = K(w � w0)2 was chosen to keep a planar group

in the same plane. The potential parameters used in these

simulations were obtained from the Gromacs force field

database and are listed in Table 2.

Table 1 Physical properties of the dicarboxylic acids studied in this work

Dicarboxylic acid Chemical formula Molar mass/g mol�1 Density/g cm�3 2,11,14Solubility in water/g per 100 g water2,11

Malonic acid (C3) HOOC–(CH2)–COOH 104.06 1.631 161Succinic acid (C4) HOOC–(CH2)2–COOH 118.09 1.572 8.8Glutaric acid (C5) HOOC–(CH2)3–COOH 132.12 1.424 116Adipic acid (C6) HOOC–(CH2)4–COOH 146.14 1.360 2.5Pimelic acid (C7) HOOC–(CH2)5–COOH 160.17 1.281 71Suberic acid (C8) HOOC–(CH2)6–COOH 174.20 1.272 0Azelaic acid (C9) HOOC–(CH2)7–COOH 188.22 1.251 0

Fig. 1 Structure of a dicarboxylic acid (C6).

Page 4: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 9377

3 Equilibration procedures

After attaching a monolayer dicarboxylic acid to the surface of

the water droplet, an energy minimization run was performed

using the conjugate gradient (CG) algorithm to relax the initial

configuration containing highly overlapped atoms. The system

was then allowed to equilibrate at 0 K for 10 ps using constant

NVE integration with the velocity-Verlet algorithm to update

position and velocity for all atoms. A timestep of 1 fs was

typically chosen to ensure energy conservation. During the

equilibration process, the temperature was controlled by rescaling

the velocities every timestep. After the system relaxation at

0 K, the coated particle was slowly heated to 260 K over 200 ps

and then allowed to equilibrate for 400 ps. Following that, the

system was heated to 300 K and allowed to equilibrate at that

temperature for up to 8 ns. For the final step of the prepara-

tion process, the simulations were switched from a constant

temperature to a constant energy calculation to ensure that the

average system temperature did not deviate by more than 10 K.

The simulations were carried out in a constant energy environ-

ment during the data generation phase. All the simulations

were run on a Linux cluster, running in parallel on 8 processors.

Our mixture method is computationally more efficient than the

fully atomistic method yet provides sufficient accuracy for the

analysis.

Results and discussion

In this section, the results from the molecular dynamics

simulations are presented and discussed. In order to study

the effect of dicarboxylic acid chain length on the structure of

coated nanoaerosols, we computed several physical properties,

including the radial density distributions, diffusion coefficients

and radial distribution functions. In order to study water

vapor processing of the coated nanoaerosols, water molecules

were impinged on the surface with a Boltzmann energy

distribution to obtain thermal sticking coefficients.

1 Structure

The time evolution of different dicarboxylic acids coated

nanoaerosol structures were monitored during the simulation

process. Different equilibrium morphologies were observed.

For C3-coated nanoaerosol, the C3 molecules slowly migrate

from the surface to the core of the nanoaerosol. This seems to

occur in two states. In the first B1 ns, the C3 molecules

dissolve and are dispersed inside the water droplet. By 5–6 ns,

however the dispersed C3 molecules dissolve further into the

water core and assemble around the center of the nanoaerosol.

The initial and final morphologies of C3-coated nanoaerosol

are presented in Fig. 2(a).

The evolution paths of C5–C9 coated nanoaerosols are

essentially similar to each other but significantly different from

that of C3-coated nanoaerosol. As a representative system, we

present the initial and final morphologies of C8-coated nano-

aerosol in Fig. 2(b). Unlike C3 molecules, C8 molecules

cannot dissolve into the water droplet. The result is phase

separation leading to the acid molecules segregation to one

side of the droplet, and forming a multilayer structure, similar

to what might be expected for the separation of oil and water.

This behavior was observed for C5–C9 coated nanoaerosols.

Apparently, for sufficiently long molecules the combined

effects of the acid–acid interaction with the hydrocarbon

backbone interactions are more competitive than the acid–

water interaction. The equilibrated structure of C4-coated

nanoaerosol is found to intermediate between the C3 and C8

cases with some of the acid molecules dissolved into the water

core while the rest formed a separate acid cluster.

To investigate the effect of a branched chain on the struc-

ture evolution of dicarboxylic acid coated nanoaerosol, a

C9 molecule was modified to add two side CH3 groups.

Fig. 2(c) shows the initial and final structures of the system.

A schematic of the branched acid structure is also shown in

Fig. 2(c). Different from previous cases, the C9_branched acid

molecules did not dissolve into the water droplet or form a

separate acid cluster, but rather maintained the initial inverted

micelle structure. Apparently, the addition of a side group

actually stabilizes the monolayer, by constraining the molecular

conformation to a linear structure.

Table 2

(1) Nonbonded force parameters

Atom e/kcal mol�1 s/A

OW–OW 0.1553 3.166HW1–HW1 0.0 0.0HW2–HW2 0.0 0.0HO–HO 0.0 0.0OA–OA 0.2029 2.955C–C 0.0970 3.361O–O 0.4122 2.6260CH2–CH2 0.1400 3.9647

(2) Harmonic bond parameters

Bond K/kcal mol�1 A�2 r0/A

OW–HW1; OW–HW2 0.0 1.0HO–OA 375.0 1.0OA–C 450.0 1.36CQO 600.0 1.23C–CH2; CH2–CH2 400.0 1.53

(3) Harmonic angle parameters

Angle K/kcal mol�1 rad�2 y0/1

HW1–OW–HW2 0.0 109.47HO–OA–C 47.5 109.5OA–CQO 60.0 124.0OA–C–CH2 60.0 115.0OQC–CH2 60.0 121.0C–CH2–CH2; CH2–CH2–CH2 55.0 111.0

(4) Proper dihedral parameters

Dihedral K/cal mol�1 n (integer) d/1

HO–OA–C–CH2; HO–OA–CQO 4.0 2 180OA–C–CH2–CH2; OQC–CH2–CH2 0.1 6 0C–CH2–CH2–CH2; CH2–CH2–CH2–CH2 1.4 3 0

(5) Improper dihedral parameters

Improper dihedral K/cal mol�1 rad�2 w0/1

C OA CH2 O 40 0.0

Page 5: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

9378 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 This journal is c the Owner Societies 2011

2 Radial density

The density profiles of the dicarboxylic acid coated nanoaerosols

were calculated as a function of radial distance r. To calculate

density, we considered a shell of thickness dr at a distance

r from the center of mass. The density at a radial distance r is

given by the mass of all the sites in that shell divided by the

volume of the shell. Due to the lack of spherical symmetry,

the density calculation could not be performed on C4–C9

dicarboxylic acid coated nanoaerosols. The calculated density

profiles for C3 and C9_branched dicarboxylic acid coated

nanoaerosols are presented in Fig. 3.

The density profiles for the C3 acid coated nanoaerosol

show, in Fig. 3(a), the acid density distribution gradually

shifting to the inside of the aerosol. After about 5–6 ns of equili-

bration, both water and acid density profiles converge to their

equilibrium values. Curiously the water density shows a bimodal

distribution, but overall the structure would indicate that the

surface is enriched in water relative to the interior of the droplet.

On the other hand the density profiles of C9_branched acid

coated nanoaerosol (Fig. 3(b)) maintain a core-shell structure.

3 Radial distribution function

The radial distribution function (RDF), g(r), is defined as the

number of atoms a distance r from a given atom compared

with the number of atoms at the same distance in an ideal

gas. In our simulation, the RDF for each atom pair is

histogrammed into 100 bins from distance 0 to the maximum

force cutoff distance. To investigate the microstructure

changes of the dicarboxylic acid coated nanoaerosols, RDFs

between different atom pairs were calculated. The RDF of the

water oxygen–oxygen pair is a good indicator of changes in the

water structure and RDF for water oxygen and the carbon in

the acid head group, tracks the interaction between water and

acid. Finally the RDF of the acids central CH2–CH2 groups

characterizes the interaction between dicarboxylic acid

molecules.

Fig. 4(a) shows the water oxygen–oxygen RDFs for C3, C8

and C9_branched acid coated nanoaerosols after equilibra-

tion. For all three RDFs, the major peak occurs at roughly

2.8 A which is the well-known average hydrogen bond length

in water. Fig. 4(b) compares the RDFs of the central

CH2–CH2 groups for the three representative nanoaerosols

after equilibration. The RDF for the C3 case exhibits a well

resolved peak at around 5 A, while the RDFs for C8 and

C9_branched chains are much more diffuse. The reader is

reminded that in the final equilibrium stage, the coating

C3 acid molecules all dissolved into the water core. Thus, a

clear peak in the RDF indicates that the C3 acid molecules

apparently form a short-range ordered structure within the

water core. The time evolution of the RDFs of acid central

CH2–CH2 in the C3 case is presented in Fig. 5(a). The system

starts from an initial ‘‘ordered structure’’ where all the acid

molecules are pointing radially out of the water droplet, after

1 ns of equilibration, the acid molecules dissolve into the water

droplet. This leads to the decrease in the peaks of RDF at 1 ns.

As the system continually evolves, the peaks in RDFs

grow which indicates the formation of a more ordered acid

structure in water. To investigate the relative orientation

Fig. 2 Structures of dicarboxylic acid coated nanoaerosols: (a) struc-

ture of C3 coated nanoaerosol, (b) structure of C8 coated nanoaerosol,

(c) structure of C9_branched coated nanoaerosol (c).

Page 6: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 9379

between the C3 acid molecules, RDFs of different groups

within the acid molecule are calculated both for the initial

configuration and for the equilibrated structure. The results

are presented in Fig. 5(b) (initial structure) and Fig. 5(c)

(equilibrated structure). In Fig. 5(b), groups 1, 2 and 3

represent the carbon atom in the head group, CH2 group,

and the carbon atom in the tail group for a C3 molecule,

respectively. As we can see from Fig. 5(b), the RDF for group

1 atoms has the highest peak intensity, while the RDF for

group 3 atoms has the lowest peak intensity, which reflects the

initial ‘‘radially outward’’ molecular configurations of C3

molecules. After equilibration, the RDFs for group 1 atoms

and group 3 atoms almost overlap which indicates a equal

separation between the group 1 atoms and group 3 atoms. The

RDF for group 2 atoms peaks at a slightly large separation.

This suggests a pairwise parallel configuration between C3

molecules (as shown in the inset of Fig. 5(c)) with the separa-

tion between the central CH2 groups slightly larger than that

for the head and tail groups.

To investigate the phase separation occurring in long chain

dicarboxylic acids coated nanoaerosols, the RDFs of the atom

pair between the water oxygen and the carbon atom in the acid

head group for the C8 case were calculated at different stages of

equilibration. The results are presented in Fig. 6. As we can see

from the figure, starting from the initial ‘‘ordered configuration’’,

the peaks in the RDFs continuously decrease, indicating the

increased separation between water and acid molecules.

4 Phase separation mechanism

In order to study the phase separation of the long chain acids

in more detail, a low surface coverage (0.6 molecules per nm2),

C8 coated nanoaerosol was prepared and allowed to evolve

from an initial ‘inverted micelle’ configuration. The images of

structure evolution of this nanoaerosol are presented in Fig. 7.

As can be seen from Fig. 7, the phase separation process

between the acid and water molecules occurs in several stages.

In the first stage, the C8 dicarboxylic acid molecules flop

down onto the water droplet surface to maximize interaction

between the two acid groups and water (0.1 ns–1 ns). The next

stage of the phase separation process is the aggregation of the

acid molecules. Here the dicarboxylic acid surface diffuses and

aggregates to form multi-layered clusters of acid molecules

which favor acid–acid interactions (1 ns–2.5 ns). In turn these

clusters also surface diffuse to aggregate into larger clusters.

5 Discussion: hydrophobicity and the structure of dicarboxylic

acid coated aqueous aerosol

The hydrophobic effect is the tendency for non-polar molecules

and water to segregate, and is an important driving force for

Fig. 3 Radial density distributions of (a) C3 and (b) C9_branched acids coated nanoaerosols as a function of simulation time.

Page 7: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

9380 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 This journal is c the Owner Societies 2011

amphiphlies self-assembly and biological folding.25 Trying to

mix hydrophobic molecules with water requires a decrease in

hydrogen bonding, which if not energetically favored, just

results in segregation of non-polar molecules from water, and

an effective attraction between hydrophobic molecules.26 If the

hydrophobic molecule is small enough, such that the disrup-

tion of the hydrogen bond network is also small, then water

molecules can adopt orientations to go around the small

hydrophobic solute without sacrificing hydrogen bonds,

or move away from the large hydrophobic object and

form an interface around it.27 The critical length scale dividing

large hydrophobic units from small has been estimated to

be B1 nm.27

The dicarboxylic acid molecules studied in this work contain

a hydrophobic hydrocarbon chain, which can participate in

the hydrophobic interaction with water. The hydrophilic parts

of the dicarboxylic acid molecules are not directly responsible

for hydrophobic assemblies, but they can affect the arrangement

of these assemblies relative to interfaces and other structures. The

structures observed in our simulation for different dicarboxylic

acid coated nanoaerosols can be explained successfully using the

knowledge of hydrophobic interactions between the acid mole-

cules and the water molecules.

Fig. 4 RDFs for different atom pairs for C3, C8 and C9_branched

dicarboxylic acid coated nanoaerosols. (a) RDFs of water O–water O

and (b) RDFs of acid central CH2–CH2 groups.

Fig. 5 (a) Time evolution of the RDFs of acid central CH2–CH2 in

the C3 case, (b) RDFs of atom groups within acid molecules at the

initial configuration; the inset shows the grouping of atoms for a

C3 molecule, (c) RDFs of atom groups within acid molecules after the

equilibration process; the inset shows a cartoon of the pairwise parallel

configuration between C3 molecules.

Page 8: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 9381

For C3 acid coated nanoaerosol, we observed a final

structure with the surface acid molecules completely dissolved

within the water core to form an ordered cluster (see g(r) in

Fig. 4(b)) at the center of the water droplet. As a result, the

original water molecules at the droplet center were excluded,

leading to a very low water density at the center (see Fig. 3(a)).

This structure is readily understood in terms of the dependence

of hydrophobic solvation on the size of hydrophobic unit.

When the C3 acid molecules cluster together they form a

sufficiently large hydrophobic unit (i.e. large volume to surface

ratio), so that the salvation free energy of the acid cluster is

lower than the overall salvation free energy of the individual

acid molecules. By forming a nearly spherical acid cluster, the

total acid surface area exposed to the unfavorable hydrophobic

interactions with water is thus minimized. The diameter of this

acid cluster is about 2 nm which is consistent with the criteria

that above 1 nm the energetic cost of assembling hydrophobic

units is significantly more favorable than the entropic cost of

keeping them separate.25

For nanoaerosols coated with low solubility dicarboxylic

acids, we observed phase separation between acid and water at

the surface of the water droplet with the acid molecules

forming a layered aggregate. The radial distribution of the

acid aggregate reveals that this aggregate structure is less

ordered than the cluster structure formed by C3 acid molecules

in water. Since the free energetic cost of dissolving large

dicarboxylic acid molecules in water is formidable, the acid

molecules will stay on the water droplet surface. However, as

we can see from Fig. 6, due to the two hydrophilic head

groups, the first step towards acid molecule aggregation is that

each individual acid molecule folds on the surface of water

droplet. This configuration is not energetically favorable since

the hydrocarbon chain in each dicarboxylic acid molecule is

close to the water. The equilibration process is to minimize this

unfavorable interaction, by minimizing the total acid surface

area exposed to water, resulting in aggregation. Since this

aggregation process happens near the surface of water droplet,

without water medium around the acid aggregate the hydro-

phobic interaction is not as strong as that in water. Therefore,

the final structure of acid aggregate is less ordered than the

structure of the C3 acid cluster in water.

For the high surface coverage (6.6 molecules per nm2),

C9_branched acid coated nanoaerosol, with the addition of

a branched chain, the C9_branched acid molecule is hard to

fold completely on the water surface due to the geometric

constraints. Thus, the hydrophobic part of the acid molecule is

not affected by the unfavorable interaction with water. The

initial structure of C9_branched acid coated nanoaerosol

keeps intact. However, for a low surface coverage C9_branched

acid coated nanoaerosol, we do observe the phase separation

between the acid and the water molecules, since in this case the

acid molecules can fold on the surface of water droplet. Fig. 8

compares the initial and equilibrated configurations of a low

surface coverage (0.6 molecules per nm2) C9_branched acid

coated nanoaerosol. A phase separation can be clearly seen.

6 Diffusion

In order to gain a better understanding of the dicarboxylic

acid coating and how it interacts with the water droplet, we

computed the diffusion coefficients for the dicarboxylic acid

molecules and the water molecules. This is achieved by

computing the mean square displacement of the dicarboxylic

acid and water molecules in the simulation and evaluating the

diffusion coefficient using eqn (3):

@hr2ðtÞi@t

¼ 2dD ð3Þ

where hr2(t)i is the mean square displacement (MSD) of the

molecules being tracked, t is time, d is the dimension available

for diffusion (in our case 3), and D is the diffusion coefficient.

The calculated diffusion coefficients for different dicarboxylic

acid molecules are presented in Fig. 9 as a function of molecular

weight. The best-fit line of the calculated self-diffusion coefficients

shows a (molecular weight)�1 dependence, which is consistent

Fig. 6 RDFs of water O-acid head group C for C8 coated nanoaerosol

at different equilibration stages.

Fig. 7 Phase separation of low coverage (0.6 molecules per nm2), C8

acid coated nanoaerosol.

Page 9: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

9382 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 This journal is c the Owner Societies 2011

with the Rouse model28 which makes a ‘‘no entanglements’’

assumption, which is likely true for the relatively small

acids being considered here. Interestingly the same molecular

weight dependence on the diffusion coefficient applies for both

the water soluble (C3 and C4) and non-soluble (C5–C9)

molecules. The reader is reminded that the C3 acid molecules

are fully dissolved into the water core and the C4 acid

molecules are partially dissolved into the water. While for

the longer chain acid molecules, a close-packed cluster phase is

formed on the surface of the water droplet. However, in each

case the acid molecules form a separate cluster either within

the water droplet (C3, C4) or on the surface of water droplet

(C5–C9). Therefore, the acid molecule self-diffusion process is

within the ‘‘bulk’’ acid materials. Interestingly, even though

the clusters formed by the acid molecules constitute only a few

hundreds of acid molecules (about 500 acid molecules), the

calculated self-diffusion coefficient reflects only the contribu-

tion from bulk diffusion. Surface diffusion is apparently not an

important factor here.

The diffusion coefficients for water molecules in different

aerosol structures were also calculated and found to be

essentially independent of the structure, with an average water

diffusion coefficient of B3.5 � 10�5 cm2 s�1 at 300 K.

7 Water vapor processing

Finally we turn our attention to how these materials might act

in the atmosphere. The accommodation of water vapor to

liquid water surfaces plays an important role in the growth of

cloud condensation nuclei into cloud droplets. Molecular

simulations of water uptake on the vapor/liquid interface of

water have been extensively studied.10,29–33 However, there are

only a few studies that have probed the effect of organic layers

on the aqueous aerosol properties.7,8 No surprising it is widely

believed that composition, concentration, structure and surface

properties of the organic coating all play a role in the water

vapor evaporation from, and condensation to such aerosols.

While some experiments have shown that surfactant films

reduce the rate of evaporation of water,34 others show for

example that butanol coatings of up to 80% monolayer

coverage do not seem to inhibit the water vapor evaporation.35

Thus considerable uncertainty exists on the effects of organic

layers on the uptake of atmospheric trace species.

In order to obtain a better understanding of the effect of

dicarboxylic acid coating on the aerosol water vapor processing,

the water sticking (mass accommodation) coefficient a was

calculated for each coated nanoaerosol system. The sticking

coefficient describing the probability of gas molecules being

incorporated into aerosol is defined as:

a ¼ number of molecules absorbed into the aerosol

number of molecules impinging the aerosol surfaceð4Þ

An accurate determination of the water sticking coefficient is

important in predicting the nucleation and growth kinetics of

cloud droplets.

To evaluate the water sticking coefficient, we placed water

monomers outside the aerosol with a random velocity pointing

to the aerosol surface, and counted the fraction at a given

thermal velocity that would be absorbed. For each simulation,

a single water molecule was placed randomly at a distance

of 5 nm from the center of the nanoaerosol, and outside the

potential cutoff distance so that initially no force was acting

between the water monomer and the coated aerosol. The water

molecule was then given an initial velocity pointing towards

the center of the aerosol which is drawn from the Maxwell–

Boltzmann speed distribution at 300 K. The trajectory of this

water monomer was monitored for 100 ps. The water monomer

was considered trapped in the aerosol if the distance between

the monomer and the coated aerosol is smaller than the size of

the aerosol, otherwise, it was considered non-absorbed. More

than 60 such water trajectories were computed for each coated

aerosol. The water sticking coefficient was then calculated as

the ratio of number of absorbed monomers to the number of

Fig. 8 Phase separation of a low surface coverage (0.6 molecules

per nm2), C9_branched acid coated nanoaerosol.

Fig. 9 Dicarboxylic acid diffusion coefficient as a function of molecular

weight. The solid line is a curve fit to a (molecular weight)�1 dependence.

Page 10: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 9383

impinging monomers. The sticking coefficients were calculated

for all the structures.

Fig. 10 presents the results of sticking coefficients calcula-

tion for different coated structures. As we can see from

Fig. 10(a), the water sticking coefficient for C3_coated nano-

aerosol is essentially unity at all impinging velocities. Since we

have also seen that the C3 molecule moves to the core, the

surface of this droplet is essentially water and thus should

behave like a pure water droplet. Our result is consistent with

other researchers’ works31,33,36 on pure water droplets showing

a mass accommodation (condensation) coefficient of B1.

Fig. 10(b) shows the sticking coefficient calculation results

for C9_branched acid coated nanoaerosol. As we can see from

the figure, the sticking coefficient is the largest for incident

speeds around the most probable speed and smaller at both

lower and higher speeds. Compared with a pure water droplet,

the sticking coefficient for C9_branched acid coated nano-

aerosol is reduced, which indicates that the acid coating

impedes the mass transfer from the gas phase to the aerosol

phase. However, the C9_branched acid coated nanoaerosol

can still process water vapor. The sticking coefficients are as

high as 70% around the most probable speeds. This is due to

the apparent hydrophilic property of the surface of the coated

nanoaerosol since the hydrophilic tails of C9_branched acids

are all exposed to the environment. This water processing,

however does not lead to absorption within the water core.

Rather the processed water is associated with the hydrophilic

tail of the C9_branched acids, over the timer period of the

simulation. This is in contrast to our prior work on straight

chain fatty acids which were shown to process the water to

the core.

Because the longer chain dicarboxylic acid molecules are not

homogeneously dispersed over the droplet surface we counted

separately the water trajectories incident onto the water phase

from the water trajectories incident onto the acid phase. As

expected, the calculated sticking coefficient for water on water

was essentially unity. However, for the water molecules

impinging on the acid, almost all of them were reflected back

to the gas phase, leading to a sticking coefficient close to zero.

This result is consistent with the acid molecules forming a

thick hydrophobic layered structure.

Conclusions

In this work, molecular dynamics simulations were used

to probe the structure and the interfacial properties of

dicarboxylic acid coated aqueous aerosol. Low molecular

weight dicarboxylic acids of various chain lengths and water

solubility (from malonic acid to azelaic acid) were chosen to

coat a water droplet. The starting point of the coated aerosol is

an inverted micelle model. For malonic acid coated aerosol,

the original surface acid molecules dissolved into the water

core. For other nanoaerosols coated with low solubility acids,

phase separation between water and acid molecules was

observed during the equilibration process. The detailed phase

separation mechanism was investigated by monitoring the

structure evolution of a 10% surface acid covered nanoaerosol.

Water vapor accommodation showed that for the C3 acid

coated nanoaerosol, a water vapor accommodation factor of

1 was found for all incident water velocities. For longer chain

coated nanoaerosols, due to the surface phase separation, a

100% sticking probability was found for water monomer

colliding onto the water phase of the coated aerosol and an

almost 0% sticking probability was found for water monomer

colliding onto the acid phase.

Acknowledgements

The authors wish to acknowledge the support of a National

Science Foundation-NIRT grant and the National Institute of

Standards and Technology.

References

1 C. N. Cruz and S. N. Pandis, A study of the ability of puresecondary organic aerosol to act as cloud condensation nuclei,Atmos. Environ., 1997, 31(15), 2205–2214.

2 N. C. Shantz, W. R. Leaitch and P. F. Caffrey, Effect of organics oflow solubility on the growth rate of cloud droplets, J. Geophys.Res., 2003, 108(D5), 4168, DOI: 10.1029/2002JD002540.

Fig. 10 Water vapor sticking coefficient for different coated nanoaerosols.

Page 11: Citethis:Phys. Chem. Chem. Phys.,2011,13 ,93749384 PAPER … · 9376 Phys. Chem. Chem. Phys.,2011,13,93749384 This ournal is c the Owner Societies 2011 potential has been studied

9384 Phys. Chem. Chem. Phys., 2011, 13, 9374–9384 This journal is c the Owner Societies 2011

3 B. Ervens, G. Feingold, S. L. Clegg and S. M. Kreidenweis, Amodeling study of aqueous production of dicarboxylic acids: 2.Implications for cloud microphysics, J. Geophys. Res., 2004,109(D15), D15206, DOI: 10.1029/2004JD004575.

4 D. J. Donaldson, H. Tervahattu, A. F. Tuck and V. Vaida,Organic aerosols and the origin of life: An hypothesis, Origins LifeEvol. Biospheres, 2004, 34(1/2), 57–67.

5 C. M. Dobson, G. B. Ellison, A. F. Tuck and V. Vaida, Atmosphericaerosols as prebiotic chemical reactors, Proc. Natl. Acad. Sci. U. S. A.,2000, 97(22), 11864–11868.

6 D. J. Donaldson and V. Vaida, The influence of organic films at theair-aqueous boundary on atmospheric processes, Chem. Rev.,2006, 106(4), 1445–1461.

7 P. Chakraborty and M. R. Zachariah, Sticking coefficient andprocessing of water vapor on organic-coated nanoaerosols,J. Phys. Chem. A, 2008, 112(5), 966–972.

8 P. Chakraborty and M. R. Zachariah, ‘‘Effective’’ negative surfacetension: A property of coated nanoaerosols relevant to the atmosphere,J. Phys. Chem. A, 2007, 111(25), 5459–5464.

9 G. B. Ellison, A. F. Tuck and V. Vaida, Atmospheric processing oforganic aerosols, J. Geophys. Res., 1999, 104(D9), 11633–11641.

10 C. E. Kolb, R. A. Cox, J. P. D. Abbatt, M. Ammann, E. J. Davis,D. J. Donaldson, B. C. Garrett, C. George, P. T. Griffiths, D. R.Hanson, M. Kulmala, G. McFiggans, U. Poschl, I. Riipinen,M. J. Rossi, Y. Rudich, P. E. Wagner, P. M. Winkler,D. R. Worsnop and C. D. O’Dowd, An overview of current issuesin the uptake of atmospheric trace gases by aerosols and clouds,Atmos. Chem. Phys., 2010, 10(21), 10561–10605.

11 M. Hori, S. Ohta, N. Murao and S. Yamagata, Activationcapability of water soluble organic substances as CCN,J. Aerosol Sci., 2003, 34(4), 419–448.

12 K. Kawamura, R. Semere, Y. Imai, Y. Fujii and M. Hayashi,Water soluble dicarboxylic acids and related compounds inAntarctic aerosols, J. Geophys. Res., 1996, 101(D13), 18721–18728.

13 M. Z. H. Rozaini and P. Brimblecombe, The odd–even behaviourof dicarboxylic acids solubility in the atmospheric aerosols, Water,Air, Soil Pollut., 2008, 198(1–4), 65–75.

14 M. Bilde, B. Svenningsson, J. Monster and T. Rosenorn,Even–odd alternation of evaporation rates and vapor pressuresof C3–C9 dicarboxylic acid aerosols, Environ. Sci. Technol., 2003,37(7), 1371–1378.

15 T. M. Raymond and S. N. Pandis, Cloud activation of single-component organic aerosol particles, J. Geophys. Res., 2002,107(D24), 4787, DOI: 10.1029/2002JD002159.

16 S. A. Safran, Statistical Thermodynamics of Surfaces, Interfaces,and Membranes, Addison-Wesley Pub., Reading, Mass, 1994.

17 A. H. Fuhrhop and T. Y. Wang, Bolaamphiphiles, Chem. Rev.,2004, 104(6), 2901–2937.

18 C. E. Corrigan and T. Novakov, Cloud condensation nucleusactivity of organic compounds: a laboratory study, Atmos. Environ.,1999, 33(17), 2661–2668.

19 S. Plimpton, Fast parallel algorithms for short-range molecular-dynamics, J. Comput. Phys., 1995, 117(1), 1–19.

20 H. J. C. Berendsen, J. R. Grigera and T. P. Straatsma, The missingterm in effective pair potentials, J. Phys. Chem., 1987, 91(24),6269–6271.

21 H. C. Berendsen, J. P. M. Postma, W. F. van Gunsteren andJ. Hermans, Interaction Models for Water in Relation to ProteinHydration, in Intermolecular Forces, ed. B. Pullman, Reidel,Dordrecht, The Netherlands, 1981, pp. 331–342.

22 B. C. Garrett, G. K. Schenter and A. Morita, Molecular simulationsof the transport of molecules across the liquid/vapor interface ofwater, Chem. Rev., 2006, 106(4), 1355–1374.

23 J. Alejandre, D. J. Tildesley and G. A. Chapela, Molecular-dynamics simulation of the orthobaric densities and surface-tension of water, J. Chem. Phys., 1995, 102(11), 4574–4583.

24 J. P. Ryckaert, G. Ciccotti and H. J. C. Berendsen, Numerical-integration of Cartesian equations of motion of a system withconstraints-molecular-dynamics of N-alkanes, J. Comput. Phys.,1977, 23(3), 327–341.

25 D. Chandler, Interfaces and the driving force of hydrophobicassembly, Nature, 2005, 437, 640–647.

26 D. Chandler, Hydrophobicity: Two faces of water, Nature, 2002,417(6888), 491–491.

27 K. Lum, D. Chandler and J. D. Weeks, Hydrophobicity at smalland large length scales, J. Phys. Chem. B, 1999, 103(22), 4570–4577.

28 G. Strobl, The physics of polymers-concepts for understanding theirstructures and behavior, Springer, 1996.

29 G. Nagayama and T. Tsuruta, A general expression for the conden-sation coefficient based on transition state theory and moleculardynamics simulation, J. Chem. Phys., 2003, 118(3), 1392–1399.

30 L. X. Dang and B. C. Garrett, Molecular mechanism of water andammonia uptake by the liquid/vapor interface of water, Chem.Phys. Lett., 2004, 385(3–4), 309–313.

31 A. Morita, M. Sugiyama, H. Kameda, S. Koda and D. R. Hanson,Mass accommodation coefficient of water: Molecular dynamicssimulation and revised analysis of droplet train/flow reactor experi-ment, J. Phys. Chem. B, 2004, 108(26), 9111–9120.

32 R. Vacha, P. Slavicek, M. Mucha, B. J. Finlayson-Pitts andP. Jungwirth, Adsorption of atmospherically relevant gases atthe air/water interface: Free energy profiles of aqueous solvationof N-2, O-2, O-3, OH, H2O, HO2, and H2O2, J. Phys. Chem. A,2004, 108(52), 11573–11579.

33 J. Vieceli, M. Roeselova and D. J. Tobias, Accommodationcoefficients for water vapor at the air/water interface, Chem. Phys.Lett., 2004, 393(1–3), 249–255.

34 J. Buajarern, L. Mitchem and J. P. Reid, Manipulation andcharacterization of aqueous sodium dodecyl sulfate/sodiumchloride aerosol particles, J. Phys. Chem. A, 2007, 111(50),13038–13045.

35 J. R. Lawrence, S. V. Glass and G. M. Nathanson, Evaporation ofwater through butanol films at the surface of supercooled sulfuricacid, J. Phys. Chem. A, 2005, 109(33), 7449–7457.

36 T. Tsuruta and G. Nagayama, Molecular dynamics studies on thecondensation coefficient of water, J. Phys. Chem. B, 2004, 108(5),1736–1743.


Related Documents